Startseite Analytic torsion of nilmanifolds with (2, 3, 5) distributions
Artikel Open Access

Analytic torsion of nilmanifolds with (2, 3, 5) distributions

  • Stefan Haller EMAIL logo
Veröffentlicht/Copyright: 2. September 2025

Abstract

We consider generic rank two distributions on five-dimensional nilmanifolds and show that the analytic torsion of their Rumin complex coincides with the Ray-Singer torsion.

1 Introduction

The classical Ray-Singer analytic torsion [45] is a spectral invariant extracted from the de Rham complex of a closed manifold. The celebrated Cheeger-Müller theorem [7,17,18,42] asserts that this analytic torsion essentially coincides with the Reidemeister torsion, a topological invariant.

Rumin and Seshadri [54] have introduced an analytic torsion of the Rumin complex on contact manifolds [48,49,51] and showed that it coincides with the Ray-Singer torsion for three-dimensional CR Seifert manifolds. Further computations for contact spheres and lens spaces have been carried out by Kitaoka [36,37]. Recently, Albin and Quan [1] proved that the Rumin-Seshadri analytic torsion differs from the Ray-Singer torsion by the integral of a local quantity, which has yet to be identified explicitly.

An analytic torsion for the Rumin complex on more general filtered manifolds [27,50,52,53] has been proposed in [29]. This analytic torsion is only defined if the osculating algebras of the filtered manifold have pure cohomology, i.e., if the grading automorphism acts by a scalar on the Lie algebra cohomology in each degree. The latter assumption appears to be rather restrictive [29, Section 3.7]. We are only aware of three types of filtered manifolds with this property: trivially filtered manifolds giving rise to the Ray-Singer torsion, contact manifolds giving rise the Rumin-Seshadri torsion, and generic rank two distributions in dimension five which are also known as (2, 3, 5) distributions [29].

A generic rank two distributions in dimension five is a rank two subbundle D in the tangent bundle of a 5-manifold M such that Lie brackets of sections of D span a subbundle [ D , D ] of rank three, and triple brackets of sections of D span all of the tangent bundle, [ D , [ D , D ] ] = T M . These geometric structures have first been studied by Cartan [16]. The Lie bracket of vector fields induces a fiberwise Lie bracket on the associated graded bundle,

t M = T M [ D , D ] [ D , D ] D D .

This is a locally trivial bundle of graded Lie algebras over M called the bundle of osculating algebras. Its fibers are all isomorphic to the five-dimensional graded nilpotent Lie algebra

g = g 3 g 2 g 1

with graded basis X 1 , X 2 g 1 , X 3 g 2 , X 4 , X 5 g 3 , and brackets

(1.1) [ X 1 , X 2 ] = X 3 , [ X 1 , X 3 ] = X 4 , [ X 2 , X 3 ] = X 5 .

The simply connected Lie group with Lie algebra g will be denoted by G . The left invariant two-plane field spanned by X 1 and X 2 provides a basic example of a (2, 3, 5) distribution on G . Unlike contact or Engel structures, (2, 3, 5) distributions do have local geometry. A distribution of this type is locally diffeomorphic to the aforementioned left invariant distribution on G if and only if Cartan’s [16] harmonic curvature tensor, a section of S 4 D * , vanishes. This curvature tensor is constructed using an equivalent description of (2, 3, 5) distributions as regular normal parabolic geometries of type ( G 2 , P ) , where G 2 denotes the split real form of the exceptional Lie group and P denotes the parabolic subgroup corresponding to the longer root, see [16,56] or [15, Section 4.3.2]. Generic rank two distributions in dimension five have attracted quite some attention recently, cf. [2,3,8,9,1214,21,2832,38,43,5558].

The Rumin complex associated with a (2, 3, 5) distribution on a five-manifold M is a natural sequence of higher order differential operators:

Γ ( q ( t M ) ) D q Γ ( q + 1 ( t M ) ) ,

where q ( t M ) denotes the vector bundle obtained by passing to the fiberwise Lie algebra cohomology of t M with trivial coefficients. The Betti numbers are rk q ( t M ) = dim H q ( g ) = 1 , 2, 3, 2, 1 for q = 0 , , 4 and the Heisenberg order of the Rumin differential D q is k q = 1 , 3, 2, 3, 1 for q = 0 , , 4 , see [11, Section 5] and [22, Example 4.21]. The Rumin differentials form a complex, D q + 1 D q = 0 , that computes the de Rham cohomology of M . Actually, there exist injective differential operators L q : Γ ( q ( t M ) ) Ω q ( M ) embedding the Rumin complex as a subcomplex in the de Rham complex and inducing isomorphisms on cohomology. Twisting with a flat complex vector bundle F , we obtain a complex of differential operators

(1.2) Γ ( q ( t M ) F ) D q Γ ( q + 1 ( t M ) F )

computing H * ( M ; F ) , the de Rham cohomology of M with coefficients in F . Rumin has shown that the sequence (1.2) becomes exact on the level of the Heisenberg principal symbol, see [50, Theorem 3], [52, Theorem 5.2], or [22, Corollary 4.18]. Hence, the Rumin complex is a Rockland [47] complex, the analogue of an elliptic complex in the Heisenberg calculus; see [22, Section 2.3] for more details.

A fiberwise graded Euclidean inner product g on t M and a fiberwise Hermitian inner product h on F give rise to L 2 inner products on Γ ( q ( t M ) F ) , which in turn provide formal adjoints D q * of the Rumin differentials in (2). Assuming M to be closed, the operator D q * D q has an infinite dimensional kernel if q > 0 , but the remaining part of its spectrum consists of isolated positive eigenvalues with finite multiplicities only. Moreover, ( D q * D q ) s is trace class for s > 10 2 k q . The number ten appears here because this is the homogeneous dimension of the filtered manifold M . Furthermore, the function tr ( D q * D q ) s admits an analytic continuation to a meromorphic function on the entire complex plane, which is holomorphic at s = 0 , see [29, Remark 2.9]. This permits to define the zeta regularized determinant det * D q by

(1.3) log det * D q 1 2 s s = 0 tr ( D q * D q ) s .

The notation det * indicates that the zero eigenspace does not contribute, i.e., we are considering the regularized product of nonzero eigenvalues. Correspondingly, the complex powers are defined to vanish on the kernel of D q * D q . The analytic torsion τ ( M , D , F , g , h ) is the graded determinant of the Rumin complex, i.e.,

(1.4) log τ ( M , D , F , g , h ) q = 0 4 ( 1 ) q log det * D q .

By Poincaré duality, see [50, Section 2], [52, Proposition 2.8], or [29, Section 3.3], we have det * D 0 = det * D 4 and det * D 1 = det * D 3 , provided h is parallel. Hence, in this (unitary) case, the torsion may be expressed in terms of three determinants,

τ ( M , D , F , g , h ) = det * 2 D 0 det * D 2 det * 2 D 1 .

Since g has pure cohomology, the grading automorphism ϕ τ Aut gr ( g ) acts as a scalar on each cohomology, namely, by τ N q on H q ( g ) , where N q = 0 , 1 , 4 , 6 , 9 , 10 for q = 0 , , 5 . These numbers are related to the Heisenberg orders of the Rumin differentials via the equation k q = N q + 1 N q . Let κ denote a common multiple of the differential orders k 0 , , k 4 . Hence, there exist natural numbers a q such that κ = a q k q . The smallest possible choice would be the lowest common multiple κ = 6 with a q = 6 , 2 , 3 , 2 , 6 for q = 0 , , 4 . Then the Rumin-Seshadri operators

(1.5) Δ q ( D q 1 D q 1 * ) a q 1 + ( D q * D q ) a q

are all of Heisenberg order 2 κ . These operators are analytically much better behaved than D q * D q . Indeed, Δ q is a Rockland [47] operator and, thus, admits a parametrix in the Heisenberg calculus [22,40,59]. Moreover, Δ q s is trace class for s > 10 2 κ , and the zeta function tr Δ q s admits an analytic continuation to a meromorphic function on the entire complex plane, which is holomorphic at s = 0 , see [23, Corollary 2]. The analytic properties of the operator D q * D q stated in the preceding paragraph can be readily deduced from the corresponding properties of Δ q , see [29, Remark 2.9] for more details. By

(1.6) ζ M , D , F , g , h ( s ) str ( N Δ s ) q = 0 5 ( 1 ) q N q tr ( Δ q s ) ,

the analytic torsion of the Rumin complex can be expressed in the following form:

(1.7) τ ( M , D , F , g , h ) = exp 1 2 κ ζ M , D , F , g , h ( 0 ) ,

which is analogous to the formulas for the Ray-Singer torsion in [45, Definition 1.6] and the Rumin-Seshadri torsion in [54, p. 728]; see [29, Eq. (34)] for more details. The good behavior of the analytic torsion function in (1.6), for instance, with respect to variation of metric choices [29, Section 2.4], serves as motivation for using powers in (1.5) such that the operators Δ q all have the same Heisenberg order.

It turns out to be convenient to incorporate the zero eigenspaces of Δ q and consider the analytic torsion of the Rumin complex as a norm D , g , h sdet H * ( M ; F ) on the graded determinant line

sdet H * ( M ; F ) = q = 0 5 ( det H q ( M ; F ) ) ( 1 ) q .

Basic properties of this torsion have been established in [29]. In the acyclic case, it reduces to (the reciprocal of) τ ( M , D , F , g , h ) . The Ray-Singer torsion, too, is best regarded as a norm on this graded determinant line [6,7]. We will denote it by RS sdet H * ( M ; F ) . The Ray-Singer torsion does not depend on metric choices since the dimension of M is odd.

In this article we will determine the analytic torsion of the Rumin complex on nilmanifolds Γ \ G , where Γ is a lattice in G , i.e., a cocompact discrete subgroup in G . We consider any left invariant (2, 3, 5) distribution D G on G and any left invariant fiberwise graded Euclidean inner product on t G . These descend to a (2, 3, 5) distribution D Γ \ G on the nilmanifold Γ \ G and a fiberwise graded Euclidean inner product g Γ \ G on t ( Γ \ G ) . For a unitary representation ρ : Γ U ( k ) , we let h ρ denote the canonical (parallel) fiberwise Hermitian inner product on the associated flat complex vector bundle F ρ G × ρ C k over Γ \ G .

We are now in a position to state our main result.

Theorem

If χ : Γ U ( 1 ) is a nontrivial unitary character, then the twisted Rumin complex on the nilmanifold Γ \ G associated with D Γ \ G and F χ is acyclic and its analytic torsion is trivial, that is,

τ ( Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ) = 1 .

In the situation of this theorem, the Ray-Singer torsion is known to be trivial too, τ RS ( Γ \ G ; F χ ) = 1 , see Lemma 6.1. Hence, it coincides with the analytic torsion of the Rumin complex. More generally, we have:

Corollary

For any unitary representation ρ : Γ U ( k ) the analytic torsion of the Rumin complex coincides with the Ray-Singer torsion, that is,

D Γ \ G , g Γ \ G , h ρ sdet H * ( Γ \ G ; F ρ ) = RS sdet H * ( Γ \ G ; F ρ ) .

These are the first (2, 3, 5) distributions for which the analytic torsion of the Rumin complex has been computed. In [29, Theorem 1.3], a partial result of this type has been obtained by exploiting a large discrete symmetry group of the distribution D Γ \ G .

The proof given below is based on the decomposition of the Rumin complex over Γ \ G into a countable direct sum D * = ρ m ( ρ ) ρ ( D * ) of Rumin complexes in irreducible unitary representations ρ of G , denoted by ρ ( D * ) . The multiplicities m ( ρ ) of the contributing representations are known explicitly through a formula due to Howe [33] and Richardson [46] and involve counting the number of solutions to certain quadratic congruences, cf. Lemma 3.2(III), (3.13), or (4.12). The zeta function in (1.6) decomposes accordingly,

(1.8) ζ Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ( s ) = ρ m ( ρ ) ζ ρ ( s )

where ζ ρ ( s ) denotes the zeta function associated with ρ ( D * ) . The values ζ ρ ( 0 ) and ζ ρ ( 0 ) have been determined in [30] for every irreducible unitary representation ρ . Even though the sum on the right-hand side in (1.8) converges only for s > 10 2 κ , we are able to conclude ζ Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ( 0 ) = 0 via regularization, whence the theorem stated earlier. This is the subtlest part of the article at hand and builds on further properties of ζ ρ ( s ) obtained in [30]. Decomposing

ρ m ( ρ ) ζ ρ ( s ) = ζ I , Γ , χ ( s ) + ζ I II , Γ , χ ( s ) + ζ I I III , Γ , χ ( s )

according to the three types of irreducible unitary representations of G , we will use Epstein zeta functions to obtain the analytic continuation of each of the three summands, cf. (5.3), (5.8), and (5.23).

Alternatively, one could try to extend Albin and Quan’s [1] analysis of the sub-Riemannian limit to show that the torsion of the Rumin complex of a (2, 3, 5) distribution differs from the Ray-Singer torsion by the integral of a local quantity cf. [1, Corollary 3]. This, too, would immediately imply the corollary stated earlier.

The remaining part of this article is organized as follows. In Section 2, we provide an explicit description of all lattices in G . In Section 3, we use the aforementioned result due to Howe [33] and Richardson [46] to decompose the space of sections of F χ into irreducible G -representations. In Section 4, we describe the corresponding decomposition of the zeta function ζ Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ( s ) . In Section 5, building on results obtained in [30], we evaluate the derivative of this zeta function at s = 0 . In Section 6, we derive the theorem and corollary stated earlier.

2 Lattices

In this section, we provide explicit descriptions of all lattices in G .

Let X 1 , , X 5 be a graded basis of g with brackets as in (1.1). For the sake of notational simplicity, we will use this basis to identify g with R 5 , that is, a vector ( x 1 , , x 5 ) t R 5 will be identified with i = 1 5 x i X i g .

The exponential map provides a diffeomorphism exp : g G . By using the Baker-Campbell-Hausdorff formula, we find

exp x 1 x 5 exp y 1 y 5 = exp z 1 z 5 ,

where

(2.1) z = x y x 1 + y 1 x 2 + y 2 x 3 + y 3 + x 1 y 2 x 2 y 1 2 x 4 + y 4 + x 1 y 3 x 3 y 1 2 + ( x 1 y 1 ) ( x 1 y 2 x 2 y 1 ) 12 x 5 + y 5 + x 2 y 3 x 3 y 2 2 + ( x 2 y 2 ) ( x 1 y 2 x 2 y 1 ) 12 .

The center of G will be denoted by Z . Clearly, Z = exp ( z ) , where z denotes the center of g , which is spanned by X 4 , X 5 . For the group of commutators, we have [ G , G ] = exp ( [ g , g ] ) and the derived subalgebra [ g , g ] is spanned by X 3 , X 4 , X 5 .

For r N and e , f , g , h , u , v Q , we consider γ ˜ i g defined by

(2.2) γ ˜ 1 = 1 0 0 0 0 , γ ˜ 2 = 0 1 0 0 0 , γ ˜ 3 = 0 0 1 r u 2 r v 2 r , γ ˜ 4 = 0 0 0 e f , γ ˜ 5 = 0 0 0 g h ,

and let Γ denote the subgroup of G generated by the exponentials:

γ i exp γ ˜ i , i = 1 , , 5 .

Let Γ denote the lattice in R 2 generated by

(2.3) 1 r 0 , 0 1 r , u 1 2 v 1 2 , e f , g h .

The following generalizes [29, Lemma 4.4].

Lemma 2.1

The subgroup Γ is a lattice in G. Moreover,

(2.4) log Γ = x 1 x 5 g x 1 , x 2 Z x 3 x 1 x 2 2 1 r Z x 4 x 1 2 x 2 12 x 1 + u 2 ( x 3 x 1 x 2 2 ) x 5 + x 1 x 2 2 12 + x 2 v 2 ( x 3 x 1 x 2 2 ) Γ

(2.5) log ( Γ [ G , G ] ) = x 1 x 5 g x 1 , x 2 = 0 x 3 1 r Z x 4 u 2 x 3 x 5 v 2 x 3 Γ

(2.6) log ( Γ Z ) = x 1 x 5 g x 1 , x 2 , x 3 = 0 x 4 x 5 Γ

(2.7) log ( [ Γ , Γ ] ) = x 1 x 5 g x 1 , x 2 = 0 x 3 Z x 4 x 3 2 , x 5 x 3 2 1 r Z

(2.8) log ( [ Γ , Γ ] Z ) = x 1 x 5 g x 1 , x 2 , x 3 = 0 x 4 , x 5 1 r Z .

Proof

With z as in (2.1), we have

z 3 z 1 z 2 2 = x 3 x 1 x 2 2 + y 3 y 1 y 2 2 x 2 y 1 z 4 z 1 2 z 2 12 z 1 + u 2 ( z 3 z 1 z 2 2 ) = x 4 x 1 2 x 2 12 x 1 + u 2 ( x 3 x 1 x 2 2 ) + y 4 y 1 2 y 2 12 y 1 + u 2 ( y 3 y 1 y 2 2 ) y 1 ( x 3 x 1 x 2 2 ) + u 1 2 x 2 y 1 + y 1 ( y 1 + 1 ) x 2 2 z 5 + z 1 z 2 2 12 + z 2 v 2 ( z 3 z 1 z 2 2 ) = x 5 + x 1 x 2 2 12 + x 2 v 2 ( x 3 x 1 x 2 2 ) + y 5 + y 1 y 2 2 12 + y 2 v 2 ( y 3 y 1 y 2 2 ) + x 2 ( y 3 y 1 y 2 2 ) + v 1 2 x 2 y 1 x 2 ( x 2 + 1 ) y 1 2

as well as:

( x 3 ) ( x 1 ) ( x 2 ) 2 = x 3 x 1 x 2 2 x 1 x 2

( x 4 ) ( x 1 ) 2 ( x 2 ) 12 ( x 1 ) + u 2 ( x 3 ) ( x 1 ) ( x 2 ) 2 = x 4 x 1 2 x 2 12 x 1 + u 2 ( x 3 x 1 x 2 2 ) x 1 x 3 x 1 x 2 2 + u 1 2 x 1 x 2 x 1 ( x 1 1 ) x 2 2 ( x 5 ) + ( x 1 ) ( x 2 ) 2 12 + ( x 2 ) v 2 ( x 3 ) ( x 1 ) ( x 2 ) 2 = x 5 + x 1 x 2 2 12 + x 2 v 2 ( x 3 x 1 x 2 2 ) + x 2 x 3 x 1 x 2 2 + v 1 2 x 1 x 2 + x 1 x 2 ( x 2 + 1 ) 2 .

By using these relations, one readily shows that the right-hand side in (2.4) is a lattice containing Γ . Using the computations,

(2.9) log ( γ 1 k γ 2 l ) = k l k l 2 k 2 l 12 k l 2 12 , log ( γ 3 r [ γ 1 , γ 2 ] 1 ) = 0 0 0 u 1 2 v 1 2 ,

(2.10) log [ γ 1 , γ 3 ] = 0 0 0 1 r 0 , log [ γ 2 , γ 3 ] = 0 0 0 0 1 r ,

it is easy to see that this lattice is generated by γ 1 , , γ 5 . Taking also into account

(2.11) log [ γ 1 , γ 2 ] = 0 0 1 1 2 1 2 , log [ γ 1 , [ γ 1 , γ 2 ] ] = 0 0 0 1 0 , log [ γ 2 , [ γ 1 , γ 2 ] ] = 0 0 0 0 1 ,

we obtain the description of [ Γ , Γ ] in (2.7). Here, the formula

log [ exp x , exp y ] = 0 0 x 1 y 2 x 2 y 1 x 1 y 3 x 3 y 1 + ( x 1 + y 1 ) ( x 1 y 2 x 2 y 1 ) 2 x 2 y 3 x 3 y 2 + ( x 2 + y 2 ) ( x 1 y 2 x 2 y 1 ) 2

for commutators is helpful. The remaining assertions are now obvious.□

Lemma 2.2

Every lattice in G is of the form considered earlier, up to a not necessarily graded automorphism of G.

Proof

Inspecting the proof of [44, Theorem 2.21] we see that every lattice Γ in G is generated by five elements γ 1 , , γ 5 such that γ 1 , γ 2 Γ , γ 3 Γ [ G , G ] , and γ 4 , γ 5 Γ Z . By [29, Lemma 4.1], there exists a not necessarily graded automorphism of g that maps log γ 1 to γ ˜ 1 and log γ 2 to γ ˜ 2 , cf. (2.2). Up to an automorphism of G , we may thus assume γ 1 = exp γ ˜ 1 and γ 2 = exp γ ˜ 2 . As γ 3 [ G , G ] and γ 4 , γ 5 Z , they must be of the form γ i = exp γ ˜ i , where γ ˜ i , i = 3,4,5 , are as indicated in (2.2) with, a priori, real numbers r , u , v , e , f , g , h . To see that these numbers must all be rational, it suffices to observe that log ( Γ Z ) is a lattice in g 3 = R 2 , which contains the vectors in (2.3) by (2.9) and (2.10), but also contains the two unit vectors in view of (2.11). Write 1 r = s t , where s and t 1 are coprime integers. Hence, there exist integers k and l such that k s + l t = 1 . Note that γ 1 , γ 2 , γ 3 k [ γ 1 , γ 2 ] l , γ 4 , γ 5 still generate Γ for we have ( γ 3 k [ γ 1 , γ 2 ] l ) s = γ 3 mod Z in view of (2.9), and we may assume that γ 4 , γ 5 generate Γ Z . Replacing γ 3 with γ 3 k [ γ 1 , γ 2 ] l , we may thus assume that the image of log γ 3 in [ g , g ] z = g 2 = R is 1 t .□

The natural homomorphism p : G G [ G , G ] R 2 gives rise to a short exact sequence of abelian groups:

(2.12) 0 Γ [ G , G ] [ Γ , Γ ] Γ [ Γ , Γ ] p ( Γ ) 0 ,

where Γ [ G , G ] [ Γ , Γ ] is a finite abelian group and p ( Γ ) Z 2 . In particular, the sequence splits, and we obtain an isomorphism

(2.13) Γ [ Γ , Γ ] Γ [ G , G ] [ Γ , Γ ] Z 2 .

The group of unitary characters of Γ , thus, is a finite union of 2-tori,

(2.14) hom ( Γ , U ( 1 ) ) A × U ( 1 ) × U ( 1 ) ,

where A = hom Γ [ G , G ] [ Γ , Γ ] , U ( 1 ) is a finite abelian group.

To specify a character χ : Γ U ( 1 ) , it suffices to know its values on the generators, χ ( γ i ) for i = 1 , , 5 .

Lemma 2.3

Suppose χ : Γ U ( 1 ) is a unitary character, and let c be a real number such that χ ( γ 3 ) = e 2 π i c r . Then there exist integers λ 0 , μ 0 such that

(2.15) λ 0 r Z , μ 0 r Z , λ 0 u 1 2 + μ 0 v 1 2 c + Z ,

(2.16) e 2 π i ( λ 0 e + μ 0 f ) = χ ( γ 4 ) , e 2 π i ( λ 0 g + μ 0 h ) = χ ( γ 5 ) .

Proof

Since Γ Z is a lattice in Z R 2 , it admits a basis consisting of two elements. Clearly, there exists a functional α z * such that the homomorphism Z U ( 1 ) , z e 2 π i α ( log z ) coincides with χ on the aforementioned basis of Γ Z . We conclude that

(2.17) χ ( γ ) = e 2 π i α ( log γ ) ,

for all γ Γ Z . Via the identification z * = ( R 2 ) * , we have α = ( λ 0 , μ 0 ) for some real numbers λ 0 and μ 0 . Putting γ = γ 4 and γ = γ 5 in (2.17), we obtain the two equations in (2.16), respectively. Substituting γ = γ 3 r [ γ 1 , γ 2 ] 1 in (2.17) and using (2.9), we obtain the last equation in (2.15), for χ ( γ 3 r [ γ 1 , γ 2 ] 1 ) = e 2 π i c . Substituting γ = [ γ 1 , γ 3 ] and γ = [ γ 2 , γ 3 ] in (2.17) and using (2.10), we obtain the first and second equations in (2.15), respectively, for χ vanishes on commutators.□

3 Decomposition into irreducible representations

If Γ is a lattice in a simply connected nilpotent Lie group N , then L 2 ( Γ \ N ) decomposes into a countable direct sum of irreducible unitary representations of N . The multiplicities of the representations appearing in this decomposition have been studied by Moore [41]. An explicit formula for these multiplicities has been conjectured by Mostow and proved, independently, by Howe [33] and Richardson [46]. More generally, for every unitary character χ : Γ U ( 1 ) , the induced representation

L 2 ( N × χ C ) = g : N C g ( γ n ) = χ ( γ ) g ( n ) for γ Γ   and  n N , g L 2 ( Γ \ N )

decomposes into a countable direct sum of irreducible unitary representations and the multiplicities are known explicitly, cf. [33, Theorem 1] and [46, Theorem 5.3].

Lemma 3.1

(Howe, Richardson) Let N be a simply connected nilpotent Lie group, suppose Γ is a lattice in N, and let χ : Γ U ( 1 ) denote a unitary character. Then L 2 ( N × χ C ) decomposes into a countable direct sum of irreducible unitary representations of N. If an irreducible unitary representation of N appears in this decomposition, then it is induced from a rational maximal character via Kirillov’s construction. Moreover, if the rational maximal character ( f ¯ , M ) induces π , then the multiplicity of π in L 2 ( N × χ C ) coincides with the (finite) cardinality ( ( M \ N ) χ Γ ) , that is, the number of Γ orbits in

( M \ N ) χ = n M \ N ( f ¯ n , M n ) i s r a t i o n a l , a n d f ¯ n Γ M n = χ Γ M n .

Here, f ¯ n ( m ) = f ¯ ( n m n 1 ) and M n = n 1 M n for n N and m M .

Let us briefly recall some of the terminology used in the preceding lemma. To this end, let n denote the Lie algebra of N , suppose f n * , and let m denote a maximal subordinate subalgebra, i.e., a subalgebra of maximal dimension in n such that f ( [ m , m ] ) = 0 . Then

(3.1) f ¯ ( m ) e 2 π i f ( log m )

defines a unitary character on the group M exp ( m ) . According to Kirillov [19,34,35], such a character induces an irreducible unitary representation of N given by right translation on the Hilbert space

h : N C h ( m n ) = f ¯ ( m ) h ( n )  for  m M   and  n N , h L 2 ( M \ N ) .

One may always assume m to be special [46, §3] and then ( f ¯ , M ) is called a maximal character [46, §4]. The group N acts by conjugation on the set of maximal characters [46, Lemma 4.1] and the stabilizer of ( f ¯ , M ) is M , see [46, Lemma 5.1]. A maximal character ( f ¯ , M ) is called rational [46, §3] if it can be obtained from a (rational) linear functional f n * mapping log Γ into the rational numbers and m is rational with respect to the rational structure on n provided by log Γ . Note that m is rational if and only if Γ M is a lattice in M , i.e., if and only if ( Γ M ) \ M is compact. Clearly, the action of Γ preserves the subset of rational maximal characters and the integrality condition f ¯ Γ M = χ Γ M .

In the remaining part of this section, we will specialize the preceding lemma to the five-dimensional Lie group G considered earlier. All irreducible unitary representations of this Lie group have been described explicitly by Dixmier in [24, Proposition 8]. There are three types of irreducible unitary representations of G , which we now describe in terms of a graded basis X 1 , , X 5 of g satisfying (1.1). If ρ is a representation of G , we let ρ denote the corresponding infinitesimal representation of g .

  1. Scalar representations: For ( α , β ) R 2 , there is an irreducible unitary representation ρ α , β of G on C such that

    (3.2) ρ α , β ( X 1 ) = 2 π i α , ρ α , β ( X 2 ) = 2 π i β ,

    and ρ α , β ( X 3 ) = ρ α , β ( X 4 ) = ρ α , β ( X 5 ) = 0 . Via Kirillov’s construction, the functional f g * with f ( X 1 ) = α , f ( X 2 ) = β , and f ( X 3 ) = f ( X 4 ) = f ( X 5 ) = 0 induces a representation isomorphic to ρ α , β . These are precisely the irreducible representations, which factor through the abelianization G [ G , G ] R 2 .

  2. Schrödinger representations: For 0 R , there is an irreducible unitary representation ρ of G on L 2 ( R ) = L 2 ( R , d θ ) such that

    (3.3) ρ ( X 1 ) = θ , ρ ( X 2 ) = 2 π i θ , ρ ( X 3 ) = 2 π i ,

    and ρ ( X 4 ) = ρ ( X 5 ) = 0 . Via Kirillov’s construction, any functional f g * with f ( X 3 ) = and f ( X 4 ) = f ( X 5 ) = 0 induces a representation isomorphic to ρ . These are precisely the irreducible unitary representations that factor through the three-dimensional Heisenberg group H = G Z but do not factor through the abelianization G [ G , G ] .

  3. Generic representations: For real numbers λ , μ , ν with ( λ , μ ) ( 0,0 ) , there is an irreducible unitary representation ρ λ , μ , ν of G on L 2 ( R ) = L 2 ( R , d θ ) such that:

    (3.4) ρ λ , μ , ν ( X 1 ) = λ ( λ 2 + μ 2 ) 1 3 θ 2 π i μ ( λ 2 + μ 2 ) 1 3 θ 2 + ν ( λ 2 + μ 2 ) 2 3 2 ρ λ , μ , ν ( X 2 ) = μ ( λ 2 + μ 2 ) 1 3 θ + 2 π i λ ( λ 2 + μ 2 ) 1 3 θ 2 + ν ( λ 2 + μ 2 ) 2 3 2 ρ λ , μ , ν ( X 3 ) = 2 π i ( λ 2 + μ 2 ) 1 3 θ , ρ λ , μ , ν ( X 4 ) = 2 π i λ , ρ λ , μ , ν ( X 5 ) = 2 π i μ .

    This differs from the representation given in [24, equation (24)] by a conjugation with a unitary scaling on L 2 ( R ) , which we have introduced for better compatibility with the grading automorphism. Note that

    ρ λ , μ , ν ( X 3 X 3 + 2 X 1 X 5 2 X 2 X 4 ) = ( 2 π ) 2 ν .

    Any functional f g * with f ( X 4 ) = λ , f ( X 5 ) = μ , and

    (3.5) f ( X 3 ) 2 + 2 f ( X 1 ) f ( X 5 ) 2 f ( X 2 ) f ( X 4 ) = ν

    induces a representation isomorphic to ρ λ , μ , ν .

The representations listed earlier are mutually nonequivalent, and they comprise all equivalence classes of irreducible unitary representations of G . Given the convention in (3.1), it turns out to be convenient to incorporate the factors 2 π in the labeling of the representations, as indicated earlier, when considering integrality with respect to a lattice Γ generated by the exponentials of γ ˜ i as in (2.2).

Lemma 3.2

Let Γ denote the lattice spanned by γ 1 , , γ 5 , where γ i = exp γ ˜ i and γ ˜ 1 , , γ ˜ 5 are as indicated in (2.2). Suppose χ : Γ U ( 1 ) is a unitary character, and let a, b, c be real numbers such that χ ( γ 1 ) = e 2 π i a , χ ( γ 2 ) = e 2 π i b , and χ ( γ 3 ) = e 2 π i c r . Then L 2 ( G × χ C ) decomposes into a countable direct sum of irreducible unitary representations with the following multiplicities:

  1. For α , β R , the representation ρ α , β appears with multiplicity

    m ( ρ α , β ) = 1 if χ Γ [ G , G ] = 1 , α a + Z , β b + Z , a n d 0 otherwise.

    The condition χ Γ [ G , G ] = 1 is equivalent to c r Z and χ ( γ 4 ) = χ ( γ 5 ) = 1 .

  2. For 0 R , the representation ρ appears with multiplicity

    m ( ρ ) = if χ Γ Z = 1 , c + r Z , a n d 0 otherwise.

    The condition χ Γ Z = 1 is equivalent to c Z and χ ( γ 4 ) = χ ( γ 5 ) = 1 .

  3. For λ , μ , ν R with ( λ , μ ) ( 0,0 ) , the multiplicity of the representation ρ λ , μ , ν vanishes unless

    (3.6) λ r Z , μ r Z , λ u 1 2 + μ v 1 2 c + Z ,

    (3.7) e 2 π i ( λ e + μ f ) = χ ( γ 4 ) , e 2 π i ( λ g + μ h ) = χ ( γ 5 ) ,

    and

    (3.8) ν = ν 0 mod r Z

    where

    (3.9) ν 0 2 ( a μ b λ ) + λ 2 μ 2 12 d 2 + ( 2 w ( λ + μ ) + λ μ d ) 2 4

    with d gcd ( λ , μ ) and

    (3.10) w c λ u 1 2 μ v 1 2 .

    In this case, the multiplicity is

    (3.11) m ( ρ λ , μ , ν ) = k Z d r Z ν = ν 0 + r k ( r k + d ) + 2 r k w mod 2 d Z .

Proof

We specialize Lemma 3.1 to N = G . Suppose f g * and put f i = f ( X i ) .

Let us begin by considering the case f [ g , g ] = 0 , i.e., f 3 = f 4 = f 5 = 0 . Via Kirillov’s construction, such a functional induces a representation isomorphic to the scalar representation labeled ρ f 1 , f 2 in (3.2). In this case, m = g is the unique (rational) maximal subordinated subalgebra. Hence, M = G . The corresponding maximal character ( f ¯ , M ) is rational iff both f 1 and f 2 are rational numbers, cf. (2.2) and (3.1). Moreover, f ¯ Γ M = χ Γ M if and only if f ¯ ( γ i ) = χ ( γ i ) for i = 1 , , 5 . As f ¯ ( γ i ) = e 2 π i f i for i = 1 , 2 and f ¯ ( γ i ) = 1 for i = 3 , 4, 5, this is the case iff f 1 a + Z , f 2 b + Z , and 1 = χ ( γ 3 ) = χ ( γ 4 ) = χ ( γ 5 ) . As ( f ¯ , M ) is fixed under the action of G , each of these representations occurs with multiplicity one in view of Lemma 3.1. According to Lemma 2.1, the group Γ [ G , G ] is generated by γ 3 , [ γ 1 , γ 3 ] , [ γ 2 , γ 3 ] , γ 3 r [ γ 1 , γ 2 ] 1 , γ 4 , γ 5 ; see (2.9) and (2.10). Hence, the condition χ Γ [ G , G ] = 1 is equivalent to χ ( γ 3 ) = χ ( γ 4 ) = χ ( γ 5 ) = 1 .

Let us next consider the case f z = 0 and f [ g , g ] 0 , i.e., f 4 = f 5 = 0 and f 3 0 . Via Kirillov’s construction, such a functional induces a representation isomorphic to the Schrödinger representation labeled ρ f 3 in (3.3). In this case, the subspace m spanned by X 2 , X 3 , X 4 , X 5 is a rational maximal subordinated subalgebra, which is stable under the action of G . Using Lemma 2.1, we see that the group Γ M is generated by

γ 2 , γ 3 , and Γ Z .

The corresponding maximal character ( f ¯ , M ) is rational iff both f 2 and f 3 are rational numbers. Moreover, f ¯ Γ M = χ Γ M if and only if f ¯ ( γ 2 ) = χ ( γ 2 ) , f ¯ ( γ 3 ) = χ ( γ 3 ) , and f ¯ Γ Z = χ Γ Z . Equivalently, f 2 b + Z , f 3 c + r Z , and 1 = χ Γ Z . Suppose g = exp ( i = 1 5 x i X i ) . A straightforward calculation yields ( Ad g * f ) ( X 2 ) = f 2 + x 1 f 3 and ( Ad g * f ) ( X i ) = f i for i = 3,4,5 . Hence, ( f ¯ g , M g = M ) is rational and f ¯ g Γ M g = χ Γ M g iff f 2 + x 1 f 3 b + Z , f 3 c + r Z , and χ Γ Z = 1 . As integral x 1 correspond to g Γ , we conclude form Lemma 3.1 that the multiplicity is f 3 , provided f 3 c + r Z , and χ Γ Z = 1 . According to Lemma 2.1, the group Γ Z is generated by [ γ 1 , γ 3 ] , [ γ 2 , γ 3 ] , γ 3 r [ γ 1 , γ 2 ] 1 , γ 4 , γ 5 . Hence, the condition χ Γ Z = 1 is equivalent to c Z and 1 = χ ( γ 4 ) = χ ( γ 5 ) .

Let us finally turn to the case f z 0 , i.e., ( f 4 , f 5 ) ( 0,0 ) . Via Kirillov’s construction, such a functional induces a representation isomorphic to the generic representation labeled ρ λ , μ , ν in (3.4), where cf. (3.5),

(3.12) λ = f 4 , μ = f 5 , ν = f 3 2 + 2 ( f 1 f 5 f 2 f 4 ) .

In this case, the subspace m spanned by f 5 X 1 f 4 X 2 , X 3 , X 4 , X 5 is the unique maximal subordinated subalgebra. We assume that the corresponding maximal character ( f ¯ , M ) is rational, i.e., f 1 f 5 f 2 f 4 , f 3 , f 4 , f 5 are all rational. As the group Γ Z is generated by [ γ 1 , γ 3 ] , [ γ 2 , γ 3 ] , γ 3 r [ γ 1 , γ 2 ] 1 , γ 4 , γ 5 , we have f ¯ Γ Z = χ Γ Z if and only if (3.6) and (3.7) hold true, cf. (2.9), (2.10), and (2.2). We assume from now on that this is the case. In particular, f 4 , f 5 are integral, and we write d = gcd ( f 4 , f 5 ) r Z . Using Lemma 2.1 and (2.1), we see that the group Γ M is generated by

γ 1 f 5 d γ 2 f 4 d = exp f 5 d f 4 d f 4 f 5 2 d 2 f 4 f 5 2 12 d 3 f 4 2 f 5 12 d 3 , γ 3 = exp 0 0 1 r u 2 r v 2 r , and Γ Z .

Hence, f ¯ Γ M = χ Γ M if and only if (furthermore)

f 1 f 5 f 2 f 4 d f 3 f 4 f 5 2 d 2 f 4 2 f 5 2 6 d 3 = a f 5 b f 4 d mod Z

and

f 3 r + ( f 4 u + f 5 v ) 2 r = c r mod Z .

Using (3.12), (3.10), and (3.9), this is readily seen to be equivalent to

ν = ν 0 + ( f 3 + λ μ 2 d ) 2 ( w ( λ + μ ) 2 + λ μ 2 d ) 2 mod 2 d Z

and

f 3 = w ( λ + μ ) 2 mod r Z .

This is the case if and only if there exists an integer k such that

f 3 = r k + w ( λ + μ ) 2

and

ν = ν 0 + r k ( r k + 2 w ( λ + μ ) + λ μ d ) mod 2 d Z .

The latter can be replaced with the equivalent condition

ν = ν 0 + r k ( r k + d ) + 2 r k w mod 2 d Z ,

for we have

d ( λ + μ ) + λ μ d = d ( 1 λ d ) ( 1 μ d ) 2 d Z ,

as λ d and μ d cannot both be even. In particular, (3.8) must hold in this case. Recall that λ , μ , and ν are invariant under coadjoint action by G . Moreover, for g = exp ( x 1 X 1 + + x 5 X 5 ) , we have ( Ad g * f ) ( X 3 ) = f 3 + x 1 f 4 + x 2 f 5 . Hence, ( f ¯ , M ) and ( f ¯ g , M g ) lie in the same Γ -orbit iff ( Ad g * f ) ( X 3 ) f 3 is divisible by d . The formula for the multiplicity now follows from Lemma 3.1.□

For integers l , r , w , n Z with l , r 1 we define

(3.13) m ( l , r , w , n ) { k Z l Z r k ( k + l ) + 2 w k n mod 2 l Z } .

Clearly,

(3.14) m ( l , r , w , n + 2 l ) = m ( l , r , w , n )

and

(3.15) n = 1 2 l m ( l , r , w , n ) = l .

With this notation, the statement in Lemma 3.2 may be expressed in the form:

(3.16) L 2 ( G × χ C ) = ( α , β ) ( a , b ) + Z 2 ρ α , β only appears if χ Γ [ G , G ] = 1 c + r Z ρ only appears if χ Γ Z = 1 ( 0,0 ) ( λ , μ ) ( λ 0 , μ 0 ) + ( Γ ) * n Z m d r , r , w , n ρ λ , μ , ν 0 + r n .

Here λ 0 , μ 0 r Z are as in Lemma 2.3, and

(3.17) ( Γ ) * = ( l , m ) R 2 l r Z , m r Z , l u 1 2 + m v 1 2 Z l e + m f Z , l g + m h Z ( r Z ) × ( r Z )

denotes the lattice dual to the lattice Γ spanned by the vectors in (2.3). Moreover, d = gcd ( λ , μ ) , and w , ν 0 are defined in (3.9) and (3.10).

4 Decomposition of the zeta function

In this section, we use the decomposition in (3.16) to decompose the twisted Rumin complex associated with a standard (2, 3, 5) distribution on the nilmanifold Γ \ G , a standard fiberwise graded Euclidean inner product on t ( Γ \ G ) , and a unitary character χ : Γ U ( 1 ) . This yields a decomposition of the corresponding zeta function, cf. (1.8).

We continue to use a graded basis X 1 , , X 5 of g satisfying the relations in (1.1). Let D G denote the left invariant distribution on G spanned by X 1 and X 2 . Left translations provide a trivialization of the tangent bundle that induces a trivialization of the bundle of osculating algebras, t G = G × g . Passing to the fiberwise Lie algebra cohomology, we obtain a trivialization of the vector bundle q ( t G ) = G × H q ( g ) . Via this identification, the untwisted Rumin differential is a left invariant differential operator D q : C ( G ) H q ( g ) C ( G ) H q + 1 ( g ) , which may be considered as follows:

(4.1) D q U ( g ) L ( H q ( g ) , H q + 1 ( g ) ) ,

where U ( g ) denotes the universal enveloping algebra of g . With respect to a particular basis of H q ( g ) , the Rumin differentials take the following form:

D 0 = X 1 X 2 D 1 = X 112 X 13 X 4 X 111 2 X 122 2 X 5 2 X 211 2 X 4 X 222 X 221 X 23 X 5 D 2 = X 12 X 3 X 11 2 0 X 22 2 3 2 X 3 X 11 2 0 X 22 2 X 21 X 3 D 3 = X 122 + X 32 X 5 2 X 112 + 2 X 4 X 111 X 222 2 X 221 2 X 5 X 211 X 31 + X 4 D 4 = X 2 , X 1 .

Here, we are using the notation X j 1 j k = X j 1 X j k . These formulas for D q have been derived in [20, Appendix B], where they are expressed with respect to a slightly different basis, see also [5, Section 4], [22, Example 4.21], or [30, Section 3.3]. The formula for D 1 has already appeared in [4, Appendix B.7].

Suppose Γ is a lattice in G that is of the form considered in Section 2, i.e., generated by the exponentials of γ ˜ i as in (2.2). Furthermore, let χ : Γ U ( 1 ) be a unitary character. For the sections of the associated flat line bundle F χ over the nilmanifold Γ \ G , we have a canonical identification

(4.2) Γ ( F χ ) = C ( G × χ C ) f C ( G , C ) f ( γ g ) = χ ( γ ) f ( g ) for  γ Γ   and  g G .

The left invariant 2-plane field D G descends to a generic distribution of rank two on the nilmanifold Γ \ G , which we denote by D Γ \ G . The trivialization q ( t G ) = G × H q ( g ) mentioned earlier gives rise to a trivialization of the vector bundle q ( t ( Γ \ G ) ) = ( Γ \ G ) × H q ( g ) . Combining this with (4.2) we obtain a canonical identification

(4.3) Γ ( q ( t ( Γ \ G ) ) F χ ) = C ( G × χ C ) H q ( g ) .

Via this identification, the twisted Rumin differentials become operators

(4.4) D q : C ( G × χ C ) H q ( g ) C ( G × χ C ) H q + 1 ( g ) ,

which are given by the matrices in (4.1), with the universal algebra now acting in the (induced) representation C ( G × χ C ) .

Let g denote the graded Euclidean inner product on g that turns X 1 , , X 5 into an orthonormal basis. The corresponding left invariant fiberwise-graded Euclidean inner product on the bundle of osculating algebras t G = G × g descends to a fiberwise graded Euclidean inner product g Γ \ G on the bundle of osculating algebras t ( Γ \ G ) over Γ \ G . The flat line bundle F χ comes with a canonical fiberwise Hermitian inner product denoted by h χ . The metrics g Γ \ G and h χ provide an L 2 inner product on the space Γ ( q ( t ( Γ \ G ) ) F χ ) . Via the identification in (4.3), this inner product corresponds to the tensor product of the standard L 2 inner product on the induced representation C ( G × χ C ) and the Hermitian inner product on H q ( g ) induced by g .

In view of (4.3), the decomposition of the induced representation described in (3.16) provides a decomposition of the twisted Rumin complex over Γ \ G into a countable orthogonal direct sum of Rumin complexes, each associated with an irreducible unitary representations of G . For every irreducible unitary representation ρ of G , we let ρ ( D q ) denote the operator induced by (4.1) in this representation, and consider the corresponding Rumin-Seshadri operator,

(4.5) Δ ρ , q ( ρ ( D q 1 ) ρ ( D q 1 ) * ) a q 1 + ( ρ ( D q ) * ρ ( D q ) ) a q .

Moreover, we define:

(4.6) ζ ρ ( s ) str ( N Δ ρ s ) q = 0 5 ( 1 ) q N q tr Δ ρ , q s .

These zeta functions are known to converge for s sufficiently large, depending on ρ , and they admit analytic continuation to meromorphic functions on the entire complex plane which are holomorphic at s = 0 , see [30, Theorem 1]. As the Rumin complex is a Rockland complex, the operators Δ ρ , q have trivial kernel for nontrivial ρ . Clearly, Δ ρ , q vanishes if ρ is the trivial representation, and so does ζ ρ ( s ) .

Using the notation from Lemma 3.2, we define

(4.7) ζ I , Γ , χ ( s ) ( α , β ) ( a , b ) + Z 2 ζ ρ α , β ( s )

if χ Γ [ G , G ] = 1 , and ζ I , Γ , χ ( s ) 0 otherwise. Moreover,

(4.8) ζ II , Γ , χ ( s ) c + r Z ζ ρ ( s )

if χ Γ Z = 1 , and ζ II I , Γ , χ ( s ) 0 otherwise. Finally,

(4.9) ζ III , Γ , χ ( s ) ( λ , μ ) ( λ 0 , μ 0 ) + ( Γ ) * n Z m ( d r , r , w , n ) ζ ρ λ , μ , ν 0 + n r ( s ) .

Here, λ 0 , μ 0 are as in Lemma 2.3, ( Γ ) * is the dual lattice in (3.17), d = gcd ( λ , μ ) , w is given in (3.10), ν 0 is given in (3.9), and the multiplicity is defined in (3.13). Moreover, we are using the common convention to decorate the summation symbol with a prime to indicate that the summand with index zero (if any) is omitted.

Lemma 4.1

For all unitary characters χ : Γ U ( 1 ) and s > 10 2 κ , we have

(4.10) ζ Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ( s ) = ζ I , Γ , χ ( s ) + ζ I I , Γ , χ ( s ) + ζ III , Γ , χ ( s ) ,

where the left-hand side has been defined in (1.6).

Proof

Via the identification in (4.3), the decomposition in (3.16) provides a decomposition of the twisted Rumin complex D * in (4.4) into a countable orthogonal direct sum of complexes,

(4.11) D * = ρ m ( ρ ) ρ ( D * ) ,

where the sum is over all irreducible unitary representations ρ of G and the multiplicities are given in Lemma 3.2. By (1.5) and (4.5), and since ρ ( Δ q ) = Δ ρ , q , this yields

Δ q = ρ m ( ρ ) Δ ρ , q .

By using (1.6) and (4.6), we obtain

ζ Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ( s ) = ρ m ( ρ ) ζ ρ ( s ) .

Both sides converge for s > 10 2 κ in view of [23, Corollary 2]. We omit the trivial representation ρ on the right hand side because ζ ρ ( s ) vanishes identically for this ρ . By combining this with the description of the multiplicities in Lemma 3.2 and the definitions in (4.7)–(4.9), we obtain the lemma.□

Subsequently, we will analyze each of the three summands in (4.10) individually.

Remark 4.2

Let us specialize to the simple lattice Γ = Γ 0 generated by X 1 and X 2 . In this case, r = 1 , u = v = 1 , e = f = g = h = 0 , c = 0 , and w = 0 . Hence,

ζ I , Γ 0 , χ ( s ) = ( α , β ) ( a , b ) + Z 2 ζ ρ α , β ( s ) , ζ I I , Γ 0 , χ ( s ) = Z ζ ρ ( s )

and

ζ I I I , Γ 0 , χ ( s ) = ( λ , μ ) Z 2 n Z m ( d , n ) ζ ρ λ , μ , ν 0 + n ( s ) ,

where d = gcd ( λ , μ ) ,

(4.12) m ( d , n ) m ( d , 1 , 0 , n ) = { k Z d Z k ( k + d ) n mod 2 d Z } ,

and ν 0 as in (3.9) with w = 0 . For this lattice, Γ 0 [ G , G ] = [ Γ 0 , Γ 0 ] by Lemma 2.1, Γ 0 [ Γ 0 , Γ 0 ] Z 2 by (2.13), and hom ( Γ 0 , U ( 1 ) ) U ( 1 ) × U ( 1 ) according to (2.14).

5 Evaluation of the zeta function

We continue to use the notation set up in the preceding section and Lemma 3.2. In particular, Γ denotes the lattice in G spanned by the exponentials of γ ˜ i as in (2.2), χ : Γ U ( 1 ) is a unitary character, and g is a graded Euclidean inner product on g such that X 1 , , X 5 is an orthonormal basis of g . This implies, that every graded automorphism ϕ Aut gr ( g ) = GL ( g 1 ) GL 2 ( R ) that preserves g g 1 also preserves g . By naturality of the Rumin differentials, ϕ D q = D q , where the dot denotes the natural left action of Aut gr ( g ) on U ( g ) L ( H q ( g ) , H q + 1 ( g ) ) , cf. (4.1). Hence, if ρ is an irreducible unitary representation of G , then

(5.1) ζ ρ ϕ ( s ) = ζ ρ ( s )

for every graded automorphism ϕ Aut gr ( g ) preserving g g 1 . Moreover, for τ > 0 , we have

(5.2) ζ ρ ϕ τ ( s ) = τ 2 κ s ζ ρ ( s ) ,

where ϕ τ Aut gr ( g ) denotes the grading automorphism acting by τ j on g j . For more details, we refer to [30, Section 3.1].

Lemma 5.1

If χ Γ [ G , G ] = 1 , then the sum in (4.7) converges for s > 1 κ , and

(5.3) ζ I , Γ , χ ( s ) = Z Epst ( 2 κ s ; a , b ) ζ ρ 1,0 ( s ) ,

where

Z Epst ( s ; a , b ) ( α , β ) ( a , b ) + Z 2 ( α 2 + β 2 ) s 2

denotes an Epstein zeta function. In particular, ζ I , Γ , χ ( s ) extends to a meromorphic function on the entire complex plane, which has a single (simple) pole at s = 1 κ with residue π κ ζ ρ 1,0 ( 1 κ ) . Moreover, for all unitary characters χ ,

(5.4) ζ I , Γ , χ ( 0 ) = 0 , ζ I , Γ , χ ( 0 ) = 0 if χ i s n o n t r i v i a l , a n d 2 κ log 2 if χ i s t r i v i a l .

Proof

W.l.o.g. χ Γ [ G , G ] = 1 . The homogeneity in (5.2) gives

ζ ρ τ α , τ β ( s ) = τ 2 κ s ζ ρ α , β ( s ) , τ > 0 .

For ( α , β ) ( 0,0 ) , we may apply (5.1) with a graded automorphism ϕ Aut gr ( g ) that acts as a rotation on g 1 , and then use the latter homogeneity to obtain

ζ ρ α , β ( s ) = ( α 2 + β 2 ) κ s ζ ρ 1,0 ( s ) .

Summing over all ( α , β ) in the shifted lattice, we obtain (5.3), cf. (4.7).

As the scalar representations ρ α , β are one-dimensional, ζ ρ α , β ( s ) is an entire function, and (4.6) yields

(5.5) ζ ρ α , β ( 0 ) = str ( N ) = q = 0 5 ( 1 ) q N q dim H q ( g ) = 0 .

By [30, Theorem 2],

(5.6) exp 1 2 κ ζ ρ α , β ( 0 ) = 1 2 .

The Epstein [25, p. 627] zeta function Z Epst ( s ; a , b ) converges for s > 2 and extends to a meromorphic function on the entire complex plane, which has a single (simple) pole at s = 2 with residue 2 π , and

(5.7) Z Epst ( 0 ; a , b ) = 1 if  a , b Z , and 0 otherwise.

Note that a , b Z if and only if χ is trivial, cf. Lemma 3.2. Combining this with (5.3), we see that the sum in (4.7) converges for s > 1 κ , and that ζ I , Γ , χ ( s ) extends to a meromorphic function on the entire complex plane, which has a single (simple) pole at s = 1 κ of residue 2 π 2 κ ζ ρ 1,0 ( 1 κ ) . Furthermore, by combining (5.5)–(5.7) with (5.3), we obtain (5.4).□

Lemma 5.2

If χ Γ Z = 1 , then the sum in (4.8) converges for s > 2 κ and

(5.8) ζ I I , Γ , χ ( s ) = r κ s + 1 Z Epst ( κ s 1 ; c r ) ζ ρ 1 ( s ) ,

where

Z Epst ( s ; a ) n Z a + n s

denotes an Epstein zeta function. In particular, ζ I I , Γ , χ ( s ) extends to a meromorphic function on the entire complex plane, which has a simple pole at s = 2 κ with residue 2 r κ ζ ρ 1 ( 2 κ ) . If c r Z , then this is the only pole of ζ I I , Γ , χ ( s ) . If c r Z , then ζ I I , Γ , χ ( s ) has one further (simple) pole at s = 1 κ with residue res s = 1 κ ζ ρ 1 ( 1 κ ) . Moreover, for all unitary characters χ ,

(5.9) ζ I I , Γ , χ ( 0 ) = 0 and ζ I I , Γ , χ ( 0 ) = 0 .

Proof

W.l.o.g. χ Γ Z = 1 . The homogeneity in (5.2) gives

ζ ρ τ 2 ( s ) = τ 2 κ s ζ ρ ( s ) , τ > 0 .

Applying (5.1) to a graded automorphism ϕ Aut gr ( g ) that acts as an isometric reflection on g 1 , we obtain ζ ρ ( s ) = ζ ρ ( s ) . Hence,

ζ ρ ( s ) = κ s ζ ρ 1 ( s )

for all 0 R . This immediately yields (5.8), cf. (4.8). In view of [30, Theorem 6, Eqs. (200), and (202)] the function ζ ρ 1 ( s ) is meromorphic on the entire complex plane, and its poles can be only located at s = 1 2 j κ , j N 0 .

The Epstein [25, p. 620] zeta function Z Epst ( s ; a ) converges for s > 1 , it extends to a meromorphic function on the entire complex plane, which has a single (simple) pole at s = 1 of residue 2, it vanishes at s 2 N , and it satisfies

(5.10) Z Epst ( 0 ; a ) = 1 if  a Z , and 0 otherwise.

Combining this with (5.8), we see that the sum in (4.8) converges for s > 2 κ and that ζ I I , Γ , χ ( s ) extends to a meromorphic function on the entire complex plane, which has a simple pole at s = 2 κ with residue 1 r 2 κ ζ ρ 1 ( 2 κ ) . If c r Z , then the zeros of Z Epst ( κ s 1 ; c r ) cancel all the poles of ζ ρ 1 ( s ) and ζ I I , Γ , χ ( s ) is holomorphic for s 2 κ . If c r Z , then all but one pole get canceled and ζ I I , Γ , χ ( s ) has one further pole at s = 1 κ with residue res s = 1 κ ζ ρ 1 ( s ) .

From [30, Proposition 1, Theorem 3], we obtain

(5.11) ζ ρ ( 0 ) = 0 and ζ ρ ( 0 ) = 0 .

Combining this with (5.8) and the fact that ζ Epst ( s ; a ) is holomorphic at s = 1 , we obtain (5.9).□

Remark 5.3

Recall that

Z Epst ( s ; a ) = ζ Hurw ( s , a ) + ζ Hurw ( s , 1 a ) if  0 < a < 1 , and 2 ζ Riem ( s ) if  a = 0 .

Here, ζ Hurw ( s , a ) = n = 0 ( n + a ) s denotes the Hurwitz zeta function, a > 0 , and ζ Riem ( s ) = ζ Hurw ( s , 1 ) = n = 1 n s denotes the Riemann zeta function. In particular, for 0 a < 1 ,

ζ Epst ( 1 ; a ) = B 2 ( a ) ,

where B 2 ( a ) = a 2 a + 1 6 denotes the second Bernoulli polynomial. Indeed, this follows from the classical identities ζ Hurw ( 1 , a ) = 1 2 B 2 ( a ) , B 2 ( 1 a ) = B 2 ( a ) , and ζ Riem ( 1 ) = 1 12 = 1 2 B 2 ( 0 ) .

Lemma 5.4

Suppose ( 0 , 0 ) ( λ , μ ) R 2 , ν 0 R , and 0 d R . Then

n Z ζ ρ λ , μ , ν 0 + 2 d n ( s )

converges for s > 2 κ , and this function admits a meromorphic continuation to the half plane [1] s > 1 8 κ , which is holomorphic at s = 0 . Moreover,

n Z ζ ρ λ , μ , ν 0 + 2 d n ( 0 ) = 0 and n Z ζ ρ λ , μ , ν 0 + 2 d n ( 0 ) = 0 .

Proof

The homogeneity in (5.2) gives

(5.12) ζ ρ τ 3 λ , τ 3 μ , τ 4 ν ( s ) = τ 2 κ s ζ ρ λ , μ , ν ( s ) , τ > 0 .

W.l.o.g. we may thus assume λ 2 + μ 2 = 1 . Furthermore, we may assume d > 0 and d ν 0 d . From [30, Theorems 4 and 5(III)], we know

(5.13) ζ ρ λ , μ , ν ( 0 ) = 0 and ζ ρ λ , μ , ν ( 0 ) = 0

for all ν R .

For ν d , we have

(5.14) ζ ρ λ , μ , ν ( s ) = ν κ s 2 E ( s ) + ν 2 κ s + 3 2 C ( s ) + R ν , ( s ) ,

where E ( s ) , C ( s ) , and R ν , ( s ) are meromorphic functions on the entire complex plane whose poles are all contained in the set

P 3 j 4 κ : j N 0 l 2 + 1 k κ : k , l N 0 \ ( N 0 ) ,

and the estimate

(5.15) R ν , ( s ) = O ( ν 5 4 )

holds uniformly on compact subsets of { s C \ P : s 1 8 κ } and for ν d . This follows from the first estimate in [30, Corollary 1] by taking, with the notation there, N = 7 4 , σ = 1 8 κ , and ε = d . By combining (5.13)–(5.15), we see that

(5.16) E ( 0 ) = C ( 0 ) = R ν , ( 0 ) = 0 and E ( 0 ) = C ( 0 ) = R ν , ( 0 ) = 0 .

From (5.14), we obtain

(5.17) n = 1 ζ ρ λ , μ , ν 0 2 d n ( s ) = ( 2 d ) κ s 2 ζ Hurw κ s 2 ; 1 ν 0 2 d E ( s ) + ( 2 d ) 2 κ s + 3 2 ζ Hurw 2 κ s 3 2 ; 1 ν 0 2 d C ( s ) + n = 1 R ν 0 2 d n , ( s ) ,

where the last sum on the right-hand side converges uniformly on compact subsets of { s C \ P : s 1 8 κ } by the estimate in (5.15). The sums making up the Hurwitz zeta functions converge for s > 2 κ and s > 5 4 κ , respectively. Hence, n = 1 ζ ρ λ , μ , ν 0 2 d n ( s ) converges for s > 2 κ , and the right-hand side in (5.17) provides the meromorphic continuation to s > 1 8 κ . By combining this with (5.16) and the fact that the Hurwitz zeta function ζ Hurw ( s ; a ) is holomorphic at s = 0 and at s = 3 2 , we conclude

(5.18) n = 1 ζ ρ λ , μ , ν 0 2 d n ( 0 ) = 0 and n = 1 ζ ρ λ , μ , ν 0 2 d n ( 0 ) = 0 .

For ν d , we have

(5.19) ζ ρ λ , μ , ν ( s ) = ν 2 κ s + 3 2 C + ( s ) + R ν , + ( s ) ,

where C + ( s ) and R ν , + ( s ) are meromorphic functions on the entire complex plane whose poles are all contained in the set P , and the estimate

(5.20) R ν , + ( s ) = O ( ν 5 4 )

holds uniformly on compact subsets of { s C \ P : s 1 8 κ } and for ν d . This follows from the second estimate in [30, Corollary 1] by taking again, with the notation there, N = 7 4 , σ = 1 8 κ , and ε = d . By combining (5.13), (5.19), and (5.20), we see that

(5.21) C + ( 0 ) = R ν , + ( 0 ) = 0 and C + ( 0 ) = R ν , + ( 0 ) = 0 .

From (5.19), we obtain

(5.22) n = 1 ζ ρ λ , μ , ν 0 + 2 d n ( s ) = ( 2 d ) 2 κ s + 3 2 ζ Hurw 2 κ s 3 2 ; 1 + ν 0 2 d C + ( s ) + n = 1 R ν 0 + 2 d n , + ( s ) ,

where the last sum on the right-hand side converges on compact subsets of { s C \ P : s 1 8 κ } by the estimate in (5.20). As before, we conclude that n = 1 ζ ρ λ , μ , ν 0 + 2 d n ( s ) converges for s > 2 κ , and the right-hand side in (5.22) provides the meromorphic continuation to s > 1 8 κ . Using (5.21), this yields

n = 1 ζ ρ λ , μ , ν 0 + 2 d n ( 0 ) = 0 and n = 1 ζ ρ λ , μ , ν 0 + 2 d n ( 0 ) = 0 .

By combining this with (5.18) and (5.13), we obtain the lemma.□

Lemma 5.5

The sum in (4.9) converges for s > 10 2 κ , and

(5.23) ζ I I I , Γ , χ ( s ) = 1 r Z Epst ( Γ ) * 2 κ s 4 3 ; λ 0 , μ 0 f ( s ) + R ˆ ( s ) ,

where

(5.24) Z Epst ( Γ ) * ( s ; λ 0 , μ 0 ) ( λ , μ ) ( λ 0 , μ 0 ) + ( Γ ) * ( λ 2 + μ 2 ) s 2

denotes an Epstein zeta function, f ( s ) denotes a meromorphic function defined in (5.32) below, and R ˆ ( s ) is an entire function. In particular, ζ I I I , Γ , χ ( s ) extends to a meromorphic function on the entire complex plane, which has at most a simple pole at s = 10 2 κ with residue 3 π κ r Area ( R 2 Γ ) f ( 10 2 κ ) .[2] If χ Γ Z 1 , then this is the only pole of ζ I I I , Γ , χ ( s ) . If χ Γ Z = 1 , then ζ I I I , Γ , χ ( s ) has one further (simple) pole at s = 2 κ with residue 1 r res s = 2 κ f ( s ) .[3] Moreover,

(5.25) ζ I I I , Γ , χ ( 0 ) = 0 and ζ I I I , Γ , χ ( 0 ) = 0 .

Proof

There exists a graded automorphism ϕ Aut gr ( g ) preserving g and acting as a rotation on g 3 such that ρ λ , μ , ν ϕ is unitarily equivalent to ρ λ 2 + μ 2 , 0 , ν , cf. [30, Eq. (48)]. Hence, [30, Section 3.1]

(5.26) tr ( e t ρ λ , μ , ν ( Δ q ) ) = tr ( e t ρ λ 2 + μ 2 , 0 , ν ( Δ q ) ) .

For t > 0 , put

(5.27) ϑ λ , μ , ν ( t ) str ( N e t ρ λ , μ , ν ( Δ ) ) q = 0 5 ( 1 ) q N q tr ( e t ρ λ , μ , ν ( Δ q ) ) .

Using (5.26) and [30, Equation (203)], we conclude that there exist a Schwartz function k S ( R 3 ) such that

(5.28) ϑ λ , μ , ν ( t ) = t 3 4 κ λ 2 + μ 2 4 k t 2 4 κ ν 2 λ 2 + μ 2 + x 2 2 t 3 4 κ λ 2 + μ 2 4 x t 6 4 κ λ 2 + μ 2 d x .

According to [30, Eq. (188)], the function k enjoys the symmetry

(5.29) k x 1 x 2 x 3 = k x 1 x 2 x 3 .

Note that ϑ λ , μ , ν ( t ) is a Schwartz function in ν , for fixed ( λ , μ ) ( 0,0 ) and t > 0 .

By the Euler-Maclaurin summation formula, for ν 0 R and 0 d R ,

n Z ϑ λ , μ , ν 0 + 2 d n ( t ) = R ϑ λ , μ , ν 0 + 2 d ν ( t ) d ν + ( 1 ) N + 1 N ! R N ν N ϑ λ , μ , ν 0 + 2 d ν ( t ) P N ( ν ) d ν ,

where P N ( x ) = B N ( x x ) denotes the periodic extension of the N th Bernoulli polynomial restricted to the unit interval. Substituting ν 0 + 2 d ν ν in the integrals, this yields

n Z ϑ λ , μ , ν 0 + 2 d n ( t ) = 1 2 d R ϑ λ , μ , ν ( t ) d ν + ( 1 ) N + 1 ( 2 d ) N 1 N ! R N ν N ϑ λ , μ , ν ( t ) P N ν ν 0 2 d d ν .

Plugging in (5.28), we obtain

n Z ϑ λ , μ , ν 0 + 2 d n ( t ) = t 3 4 κ 2 d λ 2 + μ 2 4 R 2 k t 2 4 κ ν 2 λ 2 + μ 2 + x 2 2 t 3 4 κ λ 2 + μ 2 4 x t 6 4 κ λ 2 + μ 2 d x d ν + ( 1 ) N + 1 d N 1 t 3 4 κ t N 2 κ 2 N ! λ 2 + μ 2 4 ( λ 2 + μ 2 ) N 2 R 2 N k x 1 N t 2 4 κ ν 2 λ 2 + μ 2 + x 2 2 t 3 4 κ λ 2 + μ 2 4 x t 6 4 κ λ 2 + μ 2 P N ν ν 0 2 d d x d ν .

Substituting t 2 4 κ ν 2 λ 2 + μ 2 + x 2 2 , t 3 4 κ λ 2 + μ 2 4 x ( ν , x ) in the integrals, this yields

(5.30) n Z ϑ λ , μ , ν 0 + 2 d n ( t ) = t 4 2 κ d R 2 k ν x t 3 2 κ λ 2 + μ 2 d x d ν + ( 1 ) N + 1 d N 1 t ( N 4 ) 2 κ N ! ( λ 2 + μ 2 ) N 2 R 2 N k x 1 N ν x t 3 2 κ λ 2 + μ 2 · P N 2 t 1 2 κ λ 2 + μ 2 ν t 2 κ x 2 ν 0 2 d d x d ν .

From (4.6) and (5.27), we have

ζ ρ λ , μ , ν ( s ) = 1 Γ ( s ) 0 t s 1 ϑ λ , μ , ν ( t ) d t .

Plugging in (5.30) and substituting t 3 2 κ λ 2 + μ 2 t 3 2 κ in the integrals, we obtain

(5.31) n Z ζ ρ λ , μ , ν 0 + 2 d n ( s ) = ( λ 2 + μ 2 ) ( 2 κ s ) 3 d f ( s ) + R λ , μ , ν 0 , d , N ( s ) ,

where

(5.32) f ( s ) 1 Γ ( s ) 0 t s 1 t 4 2 κ R 2 k ν x t 3 2 κ d x d ν d t

and

(5.33) R λ , μ , ν 0 , d , N ( s ) ( 1 ) N + 1 d N 1 ( λ 2 + μ 2 ) ( 2 2 N κ s ) 3 N ! Γ ( s ) 0 t s 1 t ( N 4 ) 2 κ R 2 N k x 1 N ν x t 3 2 κ P N ( λ 2 + μ 2 ) 2 3 ( 2 t 1 2 κ ν t 2 κ x 2 ) ν 0 2 d d x d ν d t .

The integral in (5.32) converges for s > 4 2 κ and extends to a meromorphic function on the entire complex plane, which has only simple poles, and these can only be located at s = ( 2 3 j ) κ , j N 0 . Note here that the inner integral over R 2 in (5.32) results in a Schwartz function in the variable t 3 2 κ which is even according to the symmetry in (5.29). The integral in (5.33) converges for s > ( 4 N ) 2 κ , and the estimate

(5.34) R λ , μ , ν 0 , d , N ( s ) = O ( d N 1 ( λ 2 + μ 2 ) ( 2 2 N κ s ) 3 )

holds uniformly for ( λ , μ ) ( 0,0 ) , ν 0 , d 0 , and uniformly for s in compact subsets contained in s > ( 4 N ) 2 κ . Hence, n Z ζ ρ λ , μ , ν 0 + 2 d n ( s ) converges for s > 4 2 κ and (5.31) provides the meromorphic extension to the entire complex plane whose poles can only occur at s = ( 2 3 j ) κ , j N 0 . Assuming N > 4 and using Lemma 5.4, we obtain from (5.31) and (5.34)

(5.35) f ( 0 ) = R λ , μ , ν 0 , d , N ( 0 ) = 0 and f ( 0 ) = R λ , μ , ν 0 , d , N ( 0 ) = 0 .

Now suppose ( 0,0 ) ( λ , μ ) ( λ 0 , μ 0 ) + ( Γ ) * as in (4.9). In particular, λ and μ are integers divisible by r , and so is d = gcd ( λ , μ ) . Also recall the integer w defined in (3.10), and the real number ν 0 is given by (3.9). By using (3.14) and (3.15), we obtain from (5.31)

(5.36) n Z m d r , r , w , n ζ ρ λ , μ , ν 0 + r n ( s ) = ( λ 2 + μ 2 ) ( 2 κ s ) 3 r f ( s ) + R ˜ λ , μ , N ( s ) ,

where the sum on the left hand side converges for s > 4 2 κ and

R ˜ λ , μ , N ( s ) n = 1 2 d r m d r , r , w , n R λ , μ , ν 0 + r n , d , N ( s ) .

Assuming N > 4 , we obtain from (5.35)

(5.37) R ˜ λ , μ , N ( 0 ) = 0 and R ˜ λ , μ , N ( 0 ) = 0 .

From (5.34) and (3.15), using the obvious estimate d λ 2 + μ 2 , we obtain

(5.38) R ˜ λ , μ , N ( s ) = O ( ( λ 2 + μ 2 ) ( 4 N 2 κ s ) 6 ) ,

uniformly for ( 0,0 ) ( λ , μ ) ( λ 0 , μ 0 ) + ( Γ ) * , and uniformly for s in compact subsets contained in s > ( 4 N ) 2 κ . Summing over all ( λ , μ ) ( 0,0 ) in the shifted lattice, we obtain from (4.9) and (5.36),

(5.39) ζ I I I , Γ , χ ( s ) = 1 r Z Epst ( Γ ) * 2 κ s 4 3 ; λ 0 , μ 0 f ( s ) + R ˆ N ( s ) ,

where the Epstein zeta function is defined in (5.24), and

(5.40) R ˆ N ( λ , μ ) ( λ 0 , μ 0 ) + ( Γ ) * R ˜ λ , μ , N ( s ) .

The sum in (5.40) converges uniformly on compact subsets of s > ( 10 N ) 2 κ by the estimate in (5.38). Assuming N > 10 , we obtain from (5.37) and (5.40),

(5.41) R ˆ N ( 0 ) = 0 and R ˆ ( 0 ) = 0 .

As R ˆ N ( s ) is an entire function, which is independent of N , we obtain (5.23).

By using a basis of the lattice ( Γ ) * , we may write

Z Epst ( Γ ) * ( s ; λ 0 , μ 0 ) = ( α , β ) ( α 0 , β 0 ) + Z 2 ( φ ( α , β ) ) s 2 ,

where ( α 0 , β 0 ) R 2 corresponds to ( λ 0 , μ 0 ) , and φ ( α , β ) is a positive quadratic form with det φ = Area ( ( R 2 ) * ( Γ ) * ) = 1 Area ( R 2 Γ ) . Hence, this Epstein [25, p. 627] zeta function converges for s > 2 , it extends to a meromorphic function on the entire complex plane which has a single (simple) pole at s = 2 with residue 2 π det φ = 2 π Area ( R 2 Γ ) , it vanishes at s 2 N , and it satisfies

(5.42) Z Epst ( Γ ) * ( 0 ; λ 0 , μ 0 ) = 1 if  ( λ 0 , μ 0 ) Γ ,  and 0 otherwise.

Note that ( λ 0 , μ 0 ) Γ if and only if χ Γ Z = 1 by Lemmas 2.3 and 3.2(II). By combining this with (5.23), we see that the sum in (4.9) converges for s > 10 2 κ and that ζ I I I , Γ , χ ( s ) extends to a meromorphic function on the entire complex plane, which has a simple pole at s = 10 2 κ with residue 1 r 3 2 κ 2 π Area ( R 2 Γ ) f ( 10 2 κ ) . If χ Γ Z 1 , then the zeros of Z Epst ( Γ ) * ( 2 κ s 4 3 ; λ 0 , μ 0 ) cancel all the poles of f ( s ) and ζ I I I , Γ , χ ( s ) is holomorphic for s 10 2 κ . If χ Γ Z = 1 , then all but one pole get canceled and ζ I I I , Γ , χ ( s ) has one further pole at s = 2 κ with residue 1 r res s = 2 κ f ( s ) . As Z Epst ( Γ ) * ( s ; λ 0 , μ 0 ) is holomorphic at s = 4 3 , we obtain (5.25) by combining (5.39) with (5.35) and (5.41).□

Remark 5.6

It is known that ζ Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ( s ) has a single pole at s = 10 2 κ , see [29, proof of Lemma 4.2] or [26, Theorem 1.1]. Hence, in view of (4.10) the poles at s = 1 κ and s = 2 κ of ζ I , Γ , χ ( s ) and ζ I I , Γ , χ ( s ) must cancel. By Lemmas 5.1, 5.2, and 5.5, this is the case if and only if

(5.43) res s = 1 κ ζ ρ 1 ( s ) = π κ ζ ρ 1,0 ( 1 κ )

and

(5.44) res s = 2 κ f ( s ) = 2 κ ζ ρ 1 ( 2 κ ) .

Moreover, we must have

(5.45) res s = 10 2 κ ζ Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ( s ) = 3 π κ r Area ( R 2 Γ ) f ( 10 2 κ ) .

Proceeding as in [30, Section 4], one readily obtains the spectra of ρ 1,0 ( Δ q ) . For q = 0,5 , we only have the eigenvalue ( 2 π ) 2 κ with multiplicity one; for q = 1,4 , we only have the eigenvalue ( 2 π ) 2 κ with multiplicity two; and for q = 2,3 , we have the eigenvalue ( 2 π ) 2 κ with multiplicity one and the eigenvalue ( 2 π ) 2 κ 2 κ 2 with multiplicity two. Plugging this into (4.6), we obtain

(5.46) ζ ρ 1,0 ( s ) = ( 2 π ) 2 κ s 4 ( 1 2 κ s 2 ) .

In [30, Section 5], the spectra of ρ ( D q * D q ) have been determined explicitly. Using the computations in [30, Section 5.4], we see that tr ρ ( Δ q ) s has a simple pole at s = 1 κ . For q = 0,5 , the residue is 1 4 π κ , for q = 1,4 the residue is 1 2 π κ , and for q = 2,3 , the residue is 1 4 π κ + 2 2 π κ . Using (4.6), this yields

(5.47) res s = 1 κ ζ ρ ( s ) = 1 2 π κ .

Combining (5.46) with (5.47) we do indeed get (5.43). We will not attempt to verify (5.44) and (5.45) independently here.

6 Proof of the main theorem

In order to prove the theorem formulated in the introduction, note first that we may w.l.o.g. assume Γ , D Γ \ G , and g Γ \ G to be of standard form. Indeed, by Lemma 2.2, we may assume that the lattice Γ is of the form considered in Section 2, i.e., generated by the exponentials of γ ˜ i as in (2.2). Moreover, according to [29, Lemma 4.2], the torsion of the Rumin complex does not depend on the choice of D Γ \ G and g Γ \ G , as long as they are induced from a left invariant (2, 3, 5) distribution on G and a left invariant fiberwise graded Euclidean inner product on t G , respectively. Hence, we may also assume that D Γ \ G and g Γ \ G are of the form considered in Section 4, i.e., D Γ \ G is induced by the left invariant two-plane field D G on G spanned by X 1 and X 2 , and g Γ \ G is induced from a graded Euclidean inner product g on g such that X 1 , , X 5 are orthonormal. Strictly speaking, [29, Lemma 4.2] only covers graded Euclidean inner products induced from left invariant sub-Riemannian metrics on D G , but the proof there can be readily extended to the generality required here by invoking Theorem 2.11 rather than Theorem 1.1.

As the Rumin complex computes the de Rham cohomology H * ( Γ \ G ; F χ ) , one can use the Leray-Serre spectral sequence to prove the acyclicity of the Rumin complex, cf. Lemma 6.1. However, the acyclicity can also be read off the decomposition provided in Lemma 3.2. Indeed, the irreducible unitary representations appearing in this decomposition are all nontrivial, as χ is assumed to be nontrivial. Since the Rumin complex is a Rockland complex, it becomes exact in every nontrivial unitary representation of G . The acyclicity of the Rumin complex thus follows from the decomposition in (4.11).

Combining Lemma 4.1 with equations (5.4), (5.9), and (5.25) yields

ζ Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ( 0 ) = ζ I , Γ , χ ( 0 ) + ζ I I , Γ , χ ( 0 ) + ζ I I I , Γ , χ ( 0 ) = 0 .

Using (1.7), we obtain

τ ( Γ \ G , D Γ \ G , F χ , g Γ \ G , h χ ) = 1 .

This completes the proof of the theorem.

To derive the corollary, we need to know the Ray-Singer torsion of Γ \ G .

Lemma 6.1

If χ : Γ U ( 1 ) is a nontrivial character, then H * ( Γ \ G ; F χ ) = 0 and the Ray-Singer torsion is trivial, τ RS ( Γ \ G ; F χ ) = 1 .

Proof

Suppose for now that χ does not vanish on Γ [ G , G ] . Recall from [44, Section 2] or Lemma 2.1 that Γ [ G , G ] is a lattice in [ G , G ] and so is the image of Γ under the canonical homomorphism G G [ G , G ] R 2 with (abelian) fiber [ G , G ] . Hence, modding out the lattice, we obtain a fibration p : Γ \ G B with a 2-torus B T 2 as a base and with typical fiber a 3-torus, V [ G , G ] Γ [ G , G ] T 3 . Using the Künneth theorem, one readily shows that the cohomology of a torus T n R n Z n with coefficients in any nontrivial flat complex line bundle is acyclic. As χ is nontrivial on Γ [ G , G ] , we therefore have H * ( V ; F χ ) = 0 . Hence, H * ( Γ \ G ; F χ ) = 0 by the Leray-Serre spectral sequence, and τ RS ( Γ \ G ; F χ ) = 1 according to [39, Corollary 0.8].

If χ vanishes on Γ [ G , G ] , then there exists a flat complex line bundle F ˜ over B such that F χ = p * F ˜ . As χ was assumed to be nontrivial, F ˜ must be nontrivial as well. Hence, H * ( B ; F ˜ ) = 0 . Let Z denote a one-parameter subgroup in the center Z that intersects Γ nontrivially. The canonical homomorphism G G [ G , G ] factors into a sequence of homomorphisms G G Z G Z G [ G , G ] . Modding out the lattice, this yields a tower of circle bundles

(6.1) Γ \ G p 3 E 2 p 2 E 1 p 1 B ,

such that p = p 1 p 2 p 3 . Using the corresponding Gysin sequences, we obtain, successively, H * ( E 1 ; p 1 * F ˜ ) = 0 , H * ( E 2 ; p 2 * p 1 * F ˜ ) = 0 , and H * ( Γ \ G ; F χ ) = 0 for we have p 3 * p 2 * p 1 * F ˜ = p * F ˜ = F χ . Moreover, τ RS ( Γ \ G ; F χ ) = 1 according to [39, Corollary 0.9].□

Proof of the corollary formulated in the introduction

For a unitary representation ρ : Γ U ( k ) , we consider the positive real number

R ( Γ , ρ ) D Γ \ G , g Γ \ G , h ρ sdet H * ( Γ \ G ; F ρ ) RS sdet H * ( Γ \ G ; F ρ ) .

If ρ is acyclic, that is, if H * ( Γ \ G ; F ρ ) = 0 , then R ( Γ , ρ ) = τ RS ( Γ \ G ; F ρ ) τ ( Γ \ G , D Γ \ G , F ρ , g Γ \ G , h ρ ) , by the very definition. Hence, combining the theorem with Lemma 6.1, we obtain

(6.2) R ( Γ , χ ) = 1 ,

for every nontrivial character χ : Γ U ( 1 ) . According to [29, Proposition 3.18], the quantity R ( Γ , χ ) depends continuously on χ . Hence, (6.2) remains true for all unitary characters χ .

If ρ : Γ U ( k ) is irreducible, then there exists a sublattice Γ in Γ and a unitary character χ : Γ U ( 1 ) such that ρ is isomorphic to the representation of Γ induced by χ , see [10, Lemma 1]. Hence, R ( Γ , ρ ) = R ( Γ , χ ) by [29, Lemma 4.3(b)] and (6.2) yields

(6.3) R ( Γ , ρ ) = 1

for irreducible ρ . Clearly, R ( Γ , ρ 1 ρ 2 ) = R ( Γ , ρ 1 ) R ( Γ , ρ 2 ) for any two unitary representations ρ 1 and ρ 2 of Γ , see [29, Lemma 4.3(a)]. Hence, (6.3) remains true for all finite dimensional unitary representation ρ of Γ . This completes the proof of the corollary.□

Acknowledgments

The author wishes to thank an anonymous referee for numerous valuable comments.

  1. Funding information: This research was funded in whole or in part by the Austrian Science Fund (FWF) Grant DOI 10.55776/P31663. For open access purposes, the author has applied a CC BY public copyright license to any author accepted manuscript version arising from this submission.

  2. Author contribution: The author confirms the sole responsibility for the conception of the study, presented results, and manuscript preparation.

  3. Conflict of interest: Author states no conflict of interest.

  4. Data availability statement: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

[1] P. Albin and H. Quan, Sub-Riemannian limit of the differential form heat kernels of contact manifolds, Int. Math. Res. Not. IMRN 8 (2022), 5818–5881. 10.1093/imrn/rnaa270Suche in Google Scholar

[2] D. An and P. Nurowski, Twistor space for rolling bodies, Comm. Math. Phys. 326 (2014), 393–414. 10.1007/s00220-013-1839-2Suche in Google Scholar

[3] D. An and P. Nurowski, Symmetric (2, 3, 5) distributions, an interesting ODE of 7th order and Plebański metric, J. Geom. Phys. 126 (2018), 93–100. 10.1016/j.geomphys.2018.01.009Suche in Google Scholar

[4] A. Baldi, B. Franchi, N. Tchou, and M. C. Tesi, Compensated compactness for differential forms in Carnot groups and applications, Adv. Math. 223 (2010), 1555–1607. 10.1016/j.aim.2009.09.020Suche in Google Scholar

[5] A. Baldi and F. Tripaldi, Comparing three possible hypoelliptic Laplacians on the 5-dimensional Cartan group via div-curl type estimates, Adv. Nonlinear Stud. 25 (2025), 33, https://doi.org/10.1515/ans-2023-0176. Suche in Google Scholar

[6] J.-M. Bismut, H. Gillet, and C. Soulé, Analytic torsion and holomorphic determinant bundles. I. Bott-Chern forms and analytic torsion, Comm. Math. Phys. 115 (1988), 49–78. 10.1007/BF01238853Suche in Google Scholar

[7] J.-M. Bismut and W. Zhang, An extension of a theorem by Cheeger and Müller, Astérisque, vol. 205, Société Mathématique de France, Paris, 1992. Suche in Google Scholar

[8] G. Bor, L. Hernández Lamoneda, and P. Nurowski, The dancing metric, G2-symmetry and projective rolling, Trans. Amer. Math. Soc. 370 (2018), 4433–4481. 10.1090/tran/7277Suche in Google Scholar

[9] G. Bor and R. Montgomery, G2 and the rolling distribution, Enseign. Math. (2) 55 (2009), no. 1–2, 157–196. 10.4171/lem/55-1-8Suche in Google Scholar

[10] I. D. Brown, Representation of finitely generated nilpotent groups, Pacific J. Math. 45 (1973), 13–26. 10.2140/pjm.1973.45.13Suche in Google Scholar

[11] R. L. Bryant, M. Eastwood, K. Neusser, and R. Gover, Some differential complexes within and beyond parabolic geometry, Differential Geometry and Tanaka Theory (in honour of Professors Reiko Miyaoka and Keizo Yamaguchi), Adv. Stud. Pure Math. vol. 82, 2019, pp. 13–40, Preprint available at https://arxiv.org/abs/1112.2142. 10.2969/aspm/08210013Suche in Google Scholar

[12] R. L. Bryant and L. Hsu, Rigidity of integral curves of rank 2 distributions, Invent. Math. 114 (1993), 435–461. 10.1007/BF01232676Suche in Google Scholar

[13] A. Čap and K. Neusser, On automorphism groups of some types of generic distributions, Differential Geom. Appl. 27 (2009), 769–779. 10.1016/j.difgeo.2009.05.003Suche in Google Scholar

[14] A. Čap and K. Sagerschnig, On Nurowski’s conformal structure associated to a generic rank two distribution in dimension five, J. Geom. Phys. 59 (2009), 901–912. 10.1016/j.geomphys.2009.04.001Suche in Google Scholar

[15] A. Čap and J. Slovák, Parabolic geometries. I. Background and general theory, Mathematical Surveys and Monographs, vol. 154, American Mathematical Society, Providence, RI, 2009. 10.1090/surv/154Suche in Google Scholar

[16] E. Cartan, Les systèmes de Pfaff, à cinq variables et les équations aux dérivées partielles du second ordre, Ann. Sci. École Norm. Sup. 27 (1910), 109–192. 10.24033/asens.618Suche in Google Scholar

[17] J. Cheeger, Analytic torsion and Reidemeister torsion, Proc. Natl. Acad. Sci. USA 74 (1977), 2651–2654. 10.1073/pnas.74.7.2651Suche in Google Scholar PubMed PubMed Central

[18] J. Cheeger, Analytic torsion and the heat equation, Ann. Math. 109 (1979), 259–322. 10.2307/1971113Suche in Google Scholar

[19] L. J. Corwin and F. P. Greenleaf, Representations of nilpotent Lie groups and their applications. Part I. Basic theory and examples, Cambridge Studies in Advanced Mathematics, vol. 18, Cambridge University Press, Cambridge, 1990. Suche in Google Scholar

[20] S. Dave and S. Haller, Graded hypoellipticity of BGG sequences, https://arxiv.org/abs/1705.01659v2. Suche in Google Scholar

[21] S. Dave and S. Haller, On 5-manifolds admitting rank two distributions of Cartan type, Trans. Amer. Math. Soc. 371 (2019), 4911–4929. 10.1090/tran/7495Suche in Google Scholar

[22] S. Dave and S. Haller, Graded hypoellipticity of BGG sequences, Ann. Global Anal. Geom. 62 (2022), 721–789. 10.1007/s10455-022-09870-0Suche in Google Scholar PubMed PubMed Central

[23] S. Dave and S. Haller, The heat asymptotics on filtered manifolds, J. Geom. Anal. 30 (2020), 337–389. 10.1007/s12220-018-00137-4Suche in Google Scholar PubMed PubMed Central

[24] J. Dixmier, Sur les représentations unitaries des groupes de Lie nilpotents. III, Canad. J. Math. 10 (1958), 321–348. 10.4153/CJM-1958-033-5Suche in Google Scholar

[25] P. Epstein, Zur Theorie allgemeiner Zetafunctionen, (German) Math. Ann. 56 (1903), 615–644. 10.1007/BF01444309Suche in Google Scholar

[26] V. Fischer, Asymptotics and zeta functions on compact nilmanifolds, J. Math. Pures Appl. 160 (2022), no. 9, 1–28. 10.1016/j.matpur.2021.12.007Suche in Google Scholar

[27] V. Fischer and F. Tripaldi, An alternative construction of the Rumin complex on homogeneousnilpotent Lie groups, Adv. Math. 429 (2023), Paper No. 109192, 39 pp. 10.1016/j.aim.2023.109192Suche in Google Scholar

[28] A. R. Gover, R. Panai, and T. Willse, Nearly Kähler geometry and (2, 3, 5)-distributions via projective holonomy, Indiana Univ. Math. J. 66 (2017), 1351–1416. 10.1512/iumj.2017.66.6089Suche in Google Scholar

[29] S. Haller, Analytic torsion of generic rank two distributions in dimension five, J. Geom. Anal. 32 (2022), Paper No. 248, 66 pp. 10.1007/s12220-022-00987-zSuche in Google Scholar PubMed PubMed Central

[30] S. Haller, Regularized determinants of the Rumin complex in irreducible unitary representations of the (2, 3, 5) nilpotent Lie group, https://arxiv.org/abs/2309.11159v2. Suche in Google Scholar

[31] M. Hammerl and K. Sagerschnig, Conformal structures associated to generic rank 2 distributions on 5-manifolds – Characterization and Killing-field decomposition, SIGMA Symmetry Integrability Geom. Methods Appl. 5 (2009), Paper No. 81, 29 pp. 10.3842/SIGMA.2009.081Suche in Google Scholar

[32] M. Hammerl and K. Sagerschnig, The twistor spinors of generic 2- and 3-distributions, Ann. Global Anal. Geom. 39 (2011), 403–425. 10.1007/s10455-010-9240-2Suche in Google Scholar

[33] R. Howe, On Frobenius reciprocity for unipotent algebraic groups over Q, Amer. J. Math. 93 (1971), 163–172. 10.2307/2373455Suche in Google Scholar

[34] A. A. Kirillov, Unitary representations of nilpotent Lie groups, Russ. Math. Surv. 17 (1962), 53–104; English translation from Russian original Uspehi Mat. Nauk 17 (1962), 57–110. Suche in Google Scholar

[35] A. A. Kirillov, Lectures on the orbit method, Graduate Studies in Mathematics, vol. 64, American Mathematical Society, Providence, RI, 2004. 10.1090/gsm/064Suche in Google Scholar

[36] A. Kitaoka, Analytic torsions associated with the Rumin complex on contact spheres, Internat. J. Math. 31 (2020), Paper No. 2050112, 16 pp. 10.1142/S0129167X20501128Suche in Google Scholar

[37] A. Kitaoka, Ray-Singer torsion and the Rumin Laplacian on lens spaces, SIGMA Symmetry Integrability Geom. Methods Appl. 18 (2022), Paper No. 091, 16 pp. Suche in Google Scholar

[38] T. Leistner, P. Nurowski, and K. Sagerschnig, New relations between G2 geometries in dimensions 5 and 7, Internat. J. Math. 28 (2017), Paper No. 1750094, 46 pp. 10.1142/S0129167X1750094XSuche in Google Scholar

[39] W. Lück, T. Schick, and T. Thielmann, Torsion and fibrations, J. Reine Angew. Math. 498 (1998), 1–33. 10.1515/crll.1998.050Suche in Google Scholar

[40] A. Melin, Lie filtrations and pseudo-differential operators, Preprint, 1982. Suche in Google Scholar

[41] C. C. Moore, Decomposition of unitary representations defined by discrete subgroups of nilpotent groups, Ann. Math. 82 (1965), 146–182. 10.2307/1970567Suche in Google Scholar

[42] W. Müller, Analytic torsion and R-torsion of Riemannian manifolds, Adv. Math. 28 (1978), 233–305. 10.1016/0001-8708(78)90116-0Suche in Google Scholar

[43] P. Nurowski, Differential equations and conformal structures, and conformal structures, J. Geom. Phys. 55 (2005), 19–49. 10.1016/j.geomphys.2004.11.006Suche in Google Scholar

[44] M. S. Raghunathan, Discrete subgroups of Lie groups, Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 68, Springer-Verlag, New York-Heidelberg, 1972. 10.1007/978-3-642-86426-1Suche in Google Scholar

[45] D. B. Ray and I. M. Singer, R-torsion and the Laplacian on Riemannian manifolds, Adv. Math. 7 (1971), 145–210. 10.1016/0001-8708(71)90045-4Suche in Google Scholar

[46] L. F. Richardson, Decomposition of the L2-space of a general compact nilmanifold, Amer. J. Math. 93 (1971), 173–190. 10.2307/2373456Suche in Google Scholar

[47] C. Rockland, Hypoellipticity on the Heisenberg group-representation-theoretic criteria, Trans. Amer. Math. Soc. 240 (1978), 1–52. 10.1090/S0002-9947-1978-0486314-0Suche in Google Scholar

[48] M. Rumin, Un complexe de formes différentielles sur les variétés de contact, Sci. Paris Sér. I Math. 310 (1990), 401–404. Suche in Google Scholar

[49] M. Rumin, Formes différentielles sur les variétés de contact, J. Differential Geom. 39 (1994), 281–330. 10.4310/jdg/1214454873Suche in Google Scholar

[50] M. Rumin, Differential geometry on C-C spaces and application to the Novikov-Shubin numbers of nilpotent Lie groups, C. R. Acad. Sci. Paris Sér. I Math. 329 (1999), 985–990. 10.1016/S0764-4442(00)88624-3Suche in Google Scholar

[51] M. Rumin, Sub-Riemannian limit of the differential form spectrum of contact manifolds, Geom. Funct. Anal. 10 (2000), 407–452. 10.1007/s000390050013Suche in Google Scholar

[52] M. Rumin, Around heat decay on forms and relations of nilpotent Lie groups, Séminaire de Théorie Spectrale et Géométrie 19, Année 2000–2001, 123–164, Sémin. Théor. Spectr. Géom. 19 (2001), Univ. Grenoble I, Saint-Martin-daHères, (2001). 10.5802/tsg.322Suche in Google Scholar

[53] M. Rumin, An introduction to spectral and differential geometry in Carnot-Carathéodory spaces, Rend. Circ. Mat. Palermo (2) 75 (2005), 139–196. Suche in Google Scholar

[54] M. Rumin and N. Seshadri, Analytic torsions on contact manifolds, Ann. Inst. Fourier (Grenoble) 62 (2012), 727–782. 10.5802/aif.2693Suche in Google Scholar

[55] K. Sagerschnig, Split octonions and generic rank two distributions in dimension five, Arch. Math. (Brno) 42 (2006), suppl, 329–339. Suche in Google Scholar

[56] K. Sagerschnig, Weyl structures for generic rank two distributions in dimension five, PhD thesis, University of Vienna, Vienna, Austria, 2008, http://othes.univie.ac.at/2186/. Suche in Google Scholar

[57] K. Sagerschnig and T. Willse, The geometry of almost Einstein (2, 3, 5) distributions, SIGMA Symmetry Integrability Geom. Methods Appl. 13 (2017), Paper No. 4, 56 pp. Suche in Google Scholar

[58] K. Sagerschnig and T. Willse, The almost Einstein operator for (2, 3, 5) distributions, Arch. Math. (Brno) 53 (2017), 347–370. 10.5817/AM2017-5-347Suche in Google Scholar

[59] E. van Erp and R. Yuncken, A groupoid approach to pseudodifferential calculi, J. Reine Angew. Math. 756 (2019), 151–182. 10.1515/crelle-2017-0035Suche in Google Scholar

Received: 2024-10-23
Accepted: 2025-07-01
Published Online: 2025-09-02

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Heruntergeladen am 9.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/agms-2025-0028/html
Button zum nach oben scrollen